Multilayer Heterojunction Anodes for Saline Wastewater Treatment

Jul 12, 2016 - Linde + Robinson Laboratories California Institute of Technology Pasadena, California 91125, United States. Environ. Sci. Technol. , 20...
0 downloads 0 Views 891KB Size
Subscriber access provided by the University of Exeter

Article

Multi-layer hetero-junction anodes for saline wastewater treatment: Design strategies and reactive species generation mechanisms Yang Yang, Jieun Shin, Justin T. Jasper, and Michael R Hoffmann Environ. Sci. Technol., Just Accepted Manuscript • DOI: 10.1021/acs.est.6b00688 • Publication Date (Web): 12 Jul 2016 Downloaded from http://pubs.acs.org on July 17, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Environmental Science & Technology is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 31

1 2

Environmental Science & Technology

Multi-layer hetero-junction anodes for saline wastewater treatment: Design strategies and reactive species generation mechanisms

3 4

Yang Yang, Jieun Shin, Justin T. Jasper, Michael R. Hoffmann*

5 6

Linde + Robinson Laboratories

7

California Institute of Technology

8

Pasadena, California 91125, United States.

9 10 11

A manuscript submitted to Environ. Sci. Technol.

12 13 14

*Corresponding author: Email: [email protected]

ACS Paragon Plus Environment

Environmental Science & Technology

15

TOC/Abstract art

16

ACS Paragon Plus Environment

Page 2 of 31

Page 3 of 31

17

Environmental Science & Technology

ABSTRACT

18

Multilayer hetero-junction SbSn/CoTi/Ir anodes, which consist of Ir0.7Ta0.3O2 bottom layers

19

coated onto a titanium base, Co-TiO2 interlayers, and over-coated discrete Sb-SnO2 islands, were

20

prepared by spray pyrolysis. The Ir0.7Ta0.3O2 bottom layer serves as an Ohmic contact to

21

facilitate electron transfer from semiconductor layers to the Ti base. The Co-TiO2 inter-layer and

22

over-coated Sb-SnO2 islands enhance the evolution of reactive chlorine. The surficial Sb-SnO2

23

islands also serve as the reactive sites for free radical generation. Experiments coupled with

24

computational kinetic simulations show that while ·OH and Cl· are initially produced on the

25

SbSn/CoTi/Ir anode surface, the dominant radical formed in solution is the dichlorine radical

26

anion, Cl2·-. The steady-state concentration of reactive radicals is ten orders of magnitude lower

27

than that of reactive chlorine. The SbSn/CoTi/Ir anode was applied to electrochemically treat

28

human wastewater. These test results show that COD and NH4+ can be removed after 2 h of

29

electrolysis with minimal energy consumption (370 kWh/kg COD and 383 kWh/kg NH4+).

30

Although free radical species contribute to COD removal, anodes designed to enhance reactive

31

chlorine production are more effective than those designed to enhance free radical production.

ACS Paragon Plus Environment

Environmental Science & Technology

32

INTRODUCTION

33

Water scarcity has been recognized as an emerging global crisis. In order to facilitate water

34

recycling and reuse, decentralized wastewater treatment has been proposed as a supplement to

35

the conventional urban wastewater system.1 Electrochemical oxidation (EO) is usually more

36

efficient than biological treatment and is often less expensive than homogenous advanced

37

oxidation processes.2, 3 In addition, the compact design, ease of automation and small carbon

38

footprint make it an ideal candidate for decentralized wastewater treatment and reuse.2, 4-7

39

The performance of EO is determined by the electrochemical generation of reactive species,

40

which largely depends on the nature of anode materials. Non-active anodes with high over-

41

potentials for oxygen evolution reaction (OER), such as those based on SnO2, PbO2, and boron-

42

doped diamond (BDD), have been extensively investigated in the previous decades.8-13 In spite

43

of their superior current efficiency for hydroxyl radical (·OH) generation, SnO2 and PbO2 anodes

44

have poor conductivity and stability. The application of BDD anodes is hindered by their high

45

cost and complicated fabrication. Conversely, Pt-group metal oxides (e.g., RuO2 and IrO2) are

46

efficient and stable catalysts for OER, exhibiting high chlorine evolution reaction (CER) activity

47

in the presence of chloride,14 although they are typically less efficient for hydroxyl radical

48

generation. Hence the development of durable anodes with high activity for both CER and

49

radical generation is an ongoing challenge.

50

Electrolyte composition is another factor in EO performance. Previously, ·OH was

51

considered as the main contributor to organic matter removal during EO.15 Recent studies have

52

pointed out that carbonate, sulfate and phosphate radicals are also potent oxidants.4, 16 Compared

53

with these anions, chloride (Cl-) in wastewater can be more readily oxidized to reactive chlorine

54

species. Enhanced electrochemical oxidation of organic compounds observed in the presence of

ACS Paragon Plus Environment

Page 4 of 31

Page 5 of 31

Environmental Science & Technology

55

Cl- has been attributed to reaction with free chlorine (Cl2, HOCl and OCl-).17, 18 More recent

56

studies have suggested that Cl· and Cl2·- might be primarily responsible for organic compound

57

degradation.7, 19, 20 However, direct experimental evidence verifying the presence or formation

58

mechanism of these radicals during electrochemical is lacking. A quantitative description of

59

reactive species formation and reactivity in Cl- solutions during the electrochemical oxidation of

60

organic contaminants has not yet been fully elucidated.

61

In this study, versatile SbSn/CoTi/Ir hetero-junction anodes with high activity for chlorine

62

and radical generation were prepared and characterized. A combination of experimental and

63

kinetic modeling approaches were undertaken to unravel anodic reactive species generation

64

mechanisms and to model their steady-state concentrations in the electrolyte. This research

65

aimed to improve the design of hetero-junction metal oxide anodes and provide new insight into

66

the mechanism of wastewater electrolysis.

67

EXPERIMENTAL SECTION

68

Electrode preparation

69

A clean Ti base metal electrode (1 cm × 1.5 cm) was polished with sand paper and etched in

70

10% HF solution for 1 min before use. Metal oxide layers were deposited on cleaned Ti surfaces

71

by spray-pyrolysis. Aqueous metal oxide precursors were atomized with 5 psi air and sprayed

72

onto the heated (300 °C) Ti foil. The resulting oxide film was annealed at 500 °C for 10 min.

73

This procedure was repeated to reach the desired mass loading, which was followed by a final

74

annealing at 500 °C for 1 h. The Ir0.7Ta0.3O2 layer precursor contained 3.5 mM IrCl3 and 1.5 mM

75

TaCl5 in isopropanol. The TiO2 precursor contained 25 mM titanium-glycolate complex prepared

76

by the hydroxo-peroxo method.21 The dopant precursor Co(NO3)2 was added to the TiO2

77

precursor at a molar fraction of 0.1. The Sb-SnO2 precursor contained 25 mM SnCl4 and 1.24

ACS Paragon Plus Environment

Environmental Science & Technology

78

mM SbCl3 dissolved in isopropanol. Anodes with only an Ir0.7Ta0.3O2 layer are denoted as Ir for

79

simplicity. Multilayer anodes with Sb-SnO2 islands, a Co doped TiO2 layer and an Ir0.7Ta0.3O2

80

layer are denoted as SbSn/CoTi/Ir. Anodes without Sb-SnO2 doping are simply denoted as

81

CoTi/Ir (with Co-doping) or Ti/Ir (without Co-doping). The mass loadings of Ir0.7Ta0.3O2, TiO2

82

and SnO2 were 0.3, 0.5 and 1.0 mg/cm2 respectively. For select anodes denoted as

83

SbSn/CoTi/Ir*, the Ir0.7Ta0.3O2 loading was reduced to 0.05 mg/cm2. Commercially available

84

IrO2-based CER anode was purchased from Nanopac® Korea.

85

Physicochemical characterization

86

X-ray photoelectron spectroscopy (XPS) was performed using a Surface Science M-Probe

87

ESCA/XPS. Morphologies and elemental composition were obtained with a ZEISS 1550VP field

88

emission scanning electron microscope (FESEM) equipped with an Oxford X-Max SDD X-ray

89

energy-dispersive spectrometer (EDS).

90

Electrochemical characterization

91

The electrolysis cell consisted of an anode in parallel with a stainless steel cathode (2×1.5

92

cm2, 5 mm separation). For characterizations obtained in NaCl electrolyte, voltages were

93

controlled versus a Ag/AgCl/Sat. NaCl reference electrode (BASI Inc). For experiments in

94

Na2SO4 electrolyte, voltages were controlled versus a Hg/Hg2SO4 reference electrode (Gamry

95

Instruments). Electrochemical double layer capacitances (Cdl) were measured by cyclic

96

voltammetry (0.1 V window centered on the open-circuit potential) in the non-Faradaic range in

97

static 30 mM Na2SO4 solution at various scan-rate (0.005-0.8 V/s).22 Electrochemical impedance

98

spectroscopy (EIS) measurements were made in a static 30 mM NaCl electrolyte. The amplitude

99

of the sinusoidal wave was 10 mV with frequencies ranging from 0.1 Hz to 100 kHz. EIS spectra

ACS Paragon Plus Environment

Page 6 of 31

Page 7 of 31

Environmental Science & Technology

100

were fitted by considering the Helmholtz layer of the anode as a Randles circuit that includes

101

solution resistance, charge transfer resistance (Rct) and capacitance.

102

Electrolysis

103

Anodes were preconditioned in 30 mM NaCl at 25 mA/cm2 for 1 h before use. The

104

uncompensated resistance (Ru) of the cell was measured by current interruption with a 200 mA

105

current bias.7 All anodic potentials were adjusted for Ru and were reported versus the normal

106

hydrogen electrode (NHE). All electrolysis experiments were in galvanostatic mode with current

107

density of 25 or 50 mA/cm2. CER tests were conducted by galvanostatic electrolysis of 30 mM

108

NaCl solution. Samples were taken at 2 min intervals over 15 min. Total chlorine (TC) and free

109

chlorine (FC) concentrations were measured using DPD (N,N-diethyl-p-phenylenediamine)

110

reagent (Hach method 10101 and 10102). Chlorine evolution rate and current efficiency (CE)

111

were calculated as previously reported.7,

112

galvanostatically. BA was chromatographically separated using a Zorbax XDB column with 10%

113

acetonitrile and 90% 0.1% formic acid as eluent. Human wastewater was collected from the

114

public solar toilet prototype located on the California Institute of Technology campus (Pasadena,

115

CA).23 Chemical oxygen demand (COD) was determined by dichromate digestion (Hach method

116

8000). Total organic carbon (TOC) was analyzed by an Aurora TOC analyzer. Anions (Cl-, ClO3-,

117

ClO4-, NO3- and PO43- ) and cations (NH4+, Na+, Ca2+ and Mg2+) were simultaneously detected by

118

ion chromatography (ICS 2000, Dionex, USA; Ionpac AS 19 and Ionpac CS 16 columns).

119

Kinetic modeling

21

Electrolysis of benzoic acid (BA) was performed

120

Kinetic modeling of CER and radical production during NaCl electrolyte electrolysis was

121

performed using Kintecus 5.75 chemical kinetic modeling software equipped with Bader-

122

Deuflhard integrator.24 The model established in this study contained 37 reactions (Table S2).19,

ACS Paragon Plus Environment

Environmental Science & Technology

Page 8 of 31

123

25-36

124

electrolysis (pH rapidly increased to 8.5 within 1 min of electrolysis) and wastewater electrolysis.

125

Unknown rate constants were obtained by fitting the experimental data with the kinetic model.

126

RESULTS AND DISCUSSION

127

Anode Characterization

The pH was held constant at 8.5, which was typical of experimental conditions during NaCl

128

Morphologies of anodes prepared by spray pyrolysis were denser and smoother than the

129

“cracked-mud” texture typical of brush-coated anodes (Figure S1). Element mapping also

130

indicated better dispersion of sub-layer Ir, top-layer Ti and Co dopant for anodes prepared by

131

spray pyrolysis. (a)

(b)

1000

CPS

800

Sn Lα 1 Ti Kα 1 Ir Mα 1

600 400

Line scan

200 0 0.0

1.5

2.0

2.5

3.0

(d)

(e)

(f) CoTi/Ir Ti/Ir

Intensity (a.u.)

Intensity (a.u.)

Ir Ti/Ir CoTi/Ir

35 30

CoTi

25

CoTi/Ir 2.0

20 1.5

15 1.0

10 0.5

5

0.0 0

0

70 68 66 64 62 60 58 56

132

Binding energy (eV)

3.5

5 μm

CoTi/Ir CoTi

Intensity (a.u.)

1.0

-Im(Z) (kOhm)

(c)

0.5

Distance (µm)

25 μm

468 466 464 462 460 458 456 454 538 536 534 532 530 528 526 524

Binding energy (eV)

Binding energy (eV)

0

5

10

1

2

3

15

4

20

5

25

Re(Z) (kOhm)

133

Figure 1. (a) FESEM image and elemental mapping of SbSn/CoTi/Ir anode. (Green: Sn;

134

Yellow: Ti; Red: Ir). (b) Cross section image of SbSn/CoTi/Ir multilayer deposited on a

135

glass slide (inset: line scan EDS spectrum). XPS spectrum of (c) Ir 4f, (d) Ti 2p and (e) O 1s

ACS Paragon Plus Environment

Page 9 of 31

Environmental Science & Technology

136

orbitals. (f) EIS of CoTi/Ir and CoTi without Ir0.7Ta0.3Ox (inset: expanded plot from Re(Z)

137

= 0-5 kOhm).

138

Deposition of Sb-SnO2 produced isolated islands on top of the Co-TiO2 layer instead of a

139

thin film (Figure 1a). This is in agreement with previously reported morphologies of Sb-SnO2

140

anodes prepared by spray pyrolysis.37 Sb-SnO2 islands were determined to be in the range of 2-4

141

µm high and were located on top of a 0.5 µm Co-TiO2 layer overlying a 0.5 µm Ir0.7Ta0.3O2 layer

142

(Figure 1b). Overlap of the EDS signal of Ir and Ti was likely due to thermal diffusion of IrO2

143

into the TiO2 layer.

144

The CoTi/Ir anode surface was primarily composed of TiO2, as evidenced by loss of 38, 39

145

distinctive iridium oxide peaks in XPS spectra (62 and 65 eV)

146

coating layers (Figure 1c). The Ti 2p peaks of CoTi/Ir (Figure 1d) shifted to slightly lower

147

binding energies compared with that of CoTi without an IrO2 under-layer. This shift was

148

ascribed to charge transfer from IrO2 to TiO2, since IrO2 has a higher work function than TiO2.40,

149

41

150

barrier between the Co-TiO2 layer and the Ti base. Electron transfer is thus facilitated, based on

151

the observed reduction of the charge transfer resistance (Rct) of anodes containing Ir (i.e., Rct was

152

reduced from 122 kΩ for CoTi to 4 kΩ for CoTi/Ir; Figure 1f).

within TiO2 or Co-TiO2

This interaction suggests the IrO2 layer acts as an electron shuttle to overcome the Schottky

153

The properties of the TiO2 layer can be modified by metal ion doping. Co doping

154

significantly increased the fraction of oxygen vacancies (531-533 eV) versus lattice oxygen

155

(529-531 eV; Figure 1e).42 This shift reflected the weakening of the oxygen binding energies of

156

CoTi/Ir versus Ti/Ir.

157

The Ir anode exhibited an onset potential of 1.32 V at 1 mA/cm2 during linear sweep

158

voltammetry in 30 mM NaCl (Figure S2), corresponding to a 0.5 V over-potential for oxygen

ACS Paragon Plus Environment

Environmental Science & Technology

Page 10 of 31

159

evolution (0.82 V at pH = 7). This is comparable to over-potentials previously reported for nano-

160

crystalline IrO2 catalysts.43,

161

potential, deposition of Sb-SnO2 raised the onset potential to 1.38 V, which closely matches the

162

CER potential (1.36 V). The observed shift in onset potential was likely due to inhibition of OER,

163

as evidenced by a decrease in the electrochemically active surface area for OER of SbSn/CoTi/Ir

164

versus Ir anodes (Figure S2). Although the Co-TiO2 interlayer only slightly affected the OER

165

onset potential, it was crucial for inhibition of OER activity. Without a Co-TiO2 coating the

166

Ir0.7Ta0.3O2 layer had access to electrolyte through cracks among the Sb-SnO2 islands, increasing

167

the electrochemically active surface area for OER and lowering the onset potential.

44

While the TiO2 or Co-TiO2 coatings barely affected the onset

168

An 83% reduction in the mass loading of Ir0.7Ta0.3O2 (SbSn/CoTi/Ir*) resulted in a relatively

169

inactive anode, based on its high onset potential (1.56 V) and low electrochemically active

170

surface area for OER (Figure S2). In addition, the loading of Ir0.7Ta0.3O2 is crucial to overall

171

anode stability. Accelerated lifetime tests show that the lifetime of SbSn/CoTi/Ir* anode at 25

172

mA/cm2 was 720 h while that of SbSn/CoTi/Ir could be up to 4 years (Figure S3).

173

Chlorine evolution

174

Coating the Ir anode with TiO2 significantly increased CER activity and current efficiency

175

during electrolysis of 30 mM NaCl solutions (Figure 2a). The increase in CER activity resulted

176

from interaction between the top TiO2 layer and the Ir0.7Ta0.3O2 sub-layer, as only TiO2 sites

177

were exposed, and TiO2 anodes without an Ir0.7Ta0.3O2 sub-layer had no CER activity (data not

178

shown).

179 180 181

It is generally accepted that CER follows the Volmer-Heyrovsky (V-H) mechanism.45, 46 The Volmer step includes the adsorption of Cl- and the discharge of an electron:

MO x + Cl- → MO x (Cl⋅) + e -

ACS Paragon Plus Environment

(1)

Page 11 of 31

Environmental Science & Technology

182

In the Heyrovsky step, the adsorbed Cl· combines with Cl- from the bulk electrolyte and releases

183

Cl2:

MO x (Cl⋅) + Cl − → MO x + Cl 2 + e −

184 185

(2)

The recombination of two Cl·via the Volmer-Tafel reaction can also produce Cl2:47

186

2 MOx (Cl⋅) → 2 MO x + Cl 2

187

Catalysts with optimal oxygen binding energies for OER normally have high activity for CER,14

188

resulting in competition between OER and CER. However, recent density functional theory

189

(DFT) calculations48 reported that selectivity towards CER could be enhanced by a monolayer

190

TiO2 coating above RuO2, slightly increasing the energy barrier for CER, but drastically raising

191

the energy barrier for OER. In support of these calculations, the TiO2 coating applied onto IrO2,

192

which has a similar oxygen binding energy to RuO2,47 significantly increased the current

193

efficiency for chlorine production. Decreased active surface area for OER and increases in OER

194

onset potential with TiO2 over-coating (Figure S2) also supported DFT calculations.

(3)

195

At the molecular level, the desorption of Cl· (eqn 2 and 3) is considered to be the rate-

196

limiting step of CER.49 Considering the positive linear relationship between oxygen binding

197

energy and chlorine binding energy,47 lowering the oxygen binding energy by Co doping (Figure

198

1e) is likely to facilitate Cl· desorption, enhancing the CER activity of CoTi/Ir compared with

199

that of Ti/Ir.

ACS Paragon Plus Environment

Environmental Science & Technology

60

(b)

(1.61)

0.6 (1.59)

(1.51)

40

(2.0)

0.4 (1.57)

(2.5)

20

0.2

C oT i/I r C oT i/I r Sb Sb Sn Sn /Ir /C oT i/I r*

(c) 30

ClFC ClO3-

20

10

0

(d)

ClFC ClO3-

20

10

1

2

3

4

Time (h)

20

10

0

0 0

ClFC ClO3ClO4-

30

0

1

2

3

Time (h)

4

0

1

2

3

Time (h)

Sb

Sn /

Ti /Ir

er m C om

200

Ir

0

ci al

0.0

CE (%)

CER (mmol/m2/s)

(1.9)

30

Concentration (mM)

CE

Concentration (mM)

CER

0.8

Concentration (mM)

(a)

Page 12 of 31

201

Figure 2. (a) Chlorine evolution rate and current efficiency measured in 30 mM NaCl at 25

202

mA/cm2. Anodic potentials are shown as numbers above each bar. Error bars represent

203

standard deviation. Electrolysis of 30 mM NaCl with (b) CoTi/Ir, (c) SbSn/CoTi/Ir and (d)

204

SbSn/CoTi/Ir* anodes. Dotted lines represent model results.

205

As expected based on the inhibition of OER by TiO2 and Co-TiO2 coatings (Figure S2), Ti/Ir

206

and CoTi/Ir anodes exhibited increased CERs compared to Ir anodes (Figure 2a). However, the

207

lower conductivity of Sb-SnO2 islands resulted in higher operating anodic potentials. The

208

relatively inferior performances of SbSn/Ir and SbSn/CoTi/Ir* indicates that the absence of a

209

CoTi layer or a reduction in Ir mass loading was detrimental to CER activity. In general, the

210

CER activities of CoTi/Ir and SbSn/CoTi/Ir were higher that of commercial anodes (Figure 2a).

211

Electrolysis of NaCl solutions with CoTi/Ir anodes resulted in gradual loss of Cl- with

212

corresponding production of HOCl/OCl- and ClO3- (Figure 2b). Similar results were observed

213

with SbSn/CoTi/Ir anodes (Figure 2c) except with a higher production of ClO3-. Testing of

214

SbSn/CoTi/Ir* anodes showed formation of ClO4- (Figure 2d), in agreement with previous

215

studies demonstrating that non-active electrodes produce ClO4- more readily than active

216

electrode.50, 51

ACS Paragon Plus Environment

4

Page 13 of 31

217 218

219 220

Environmental Science & Technology

Kinetic modeling was performed to estimate rate constants of key reactions involved in the evolution of Cl2 (eqn 4) and the pH-dependent equilibria of Cl2, HOCl and OCl- (Table S2). k'

1 2Cl-  → Cl2

(4)

The direct oxidation of HOCl/OCl- into ClO3-,52 and ClO3- into ClO4-50 were considered as well: k'

221

2a 2 MO x+1 + OCl -  → 2 MO x + ClO3-

222

2b 2MO x+1 + HOCl  → 2MO x + ClO-3 + H +

223

3 MO x+1 + ClO3−  → MO x + ClO-4

(5)

k'

(6)

k'

224

(7)

The overall kinetics could be treated as first-order reaction in series:

225

d[Cl- ] = −k1[Cl- ] dt

(8)

226

d[FC] = k1[Cl- ] − k2' [MO x+1 ][FC] = k1[Cl- ] − k2 [FC] dt

(9)

227

d[ClO3− ] ' = k2 [MO x+1 ][FC]− k3'[MO x+1 ][ClO3− ] = k2 [FC] − k3[ClO3− ] dt

(10)

228

d[ClO-4 ] ' = k3[MO x+1 ][ClO-3 ] = k3[ClO-3 ] dt

(11)

229

FC formation rates (k1) for the CoTi/Ir and SbSn/CoTi/Ir anodes were found to be more than

230

two orders of magnitude higher than ClO3- formation rates (k2) (Table 1). ClO4- formation rates

231

(k3) were only calculated for SbSn/CoTi/Ir* anodes and were lower than FC and ClO3- formation

232

rates (k1 and k2), in line with previous research showing that the oxidation of ClO3- to ClO4- is

233

sluggish.53 Increased current density (50 vs 30 mA/cm2) with SbSn/CoTi/Ir anodes did not

234

markedly increase the FC concentration, but instead resulted in greater ClO3- production (Figure

ACS Paragon Plus Environment

Environmental Science & Technology

Page 14 of 31

235

S4). Model fitting showed that an increase in current density led to an increase in apparent rate

236

constants (Table 1), which can be explained by the Butler-Volmer formulation.54 That is, the

237

simultaneous increase of FC production rate (k1) and FC oxidation rate (k2) results in a less

238

pronounced increase of d[FC]/dt in eqn 9, which explains the inefficient chlorine accumulation

239

at 50 mA/cm2 (Figure S4).

240

It appears that an increase in the Cl- concentration increases the FC concentration more

241

efficiently than an increased current density. As expected, doubling Cl- concentrations (i.e., 60 vs.

242

30 mM) during electrolysis with CoTi/Ir, SbSn/CoTi/Ir and SbSn/CoTi/Ir* anodes at 25 mA/cm2

243

resulted in approximately double the peak FC concentration (Figure S4).

244

Table 1. Rate constants estimated by kinetic modeling. 25 mA/cm2

50 mA/cm2

CoTi/Ir SbSn/CoTi/Ir SbSn/CoTi/Ir* SbSn/CoTi/Ir k1 (10-3 M-1s-1)

2.69

7.65

3.15

14.5

k2 (10-5 M-1s-1)

1.88

6.06

32.8

19.2

k3 (10-4 M-1s-1)

-

-

2.18

0.0877

rHO· (10-8 M s-1)

-

-

1.38

0.792

kCl· (10-6 s-1)

-

-

6.48

14.9

245 246

Modeling of the electrolytic process in 60 mM NaCl gave results that were consistent with

247

the experimental data (Figure S4). However, the actual increase of ClO4- production with

248

SbSn/CoTi/Ir* anode was less than predicted. This may have been because high Cl-

249

concentrations inhibited ClO4- formation by blocking active sites for the oxidation of ClO3- to

ACS Paragon Plus Environment

Page 15 of 31

Environmental Science & Technology

250

ClO4-.51 Nevertheless, the kinetic model as presented provides a simple but powerful tool for the

251

optimization of CER.

252

Radical generation.

253

In addition to free chlorine, electrolysis of NaCl aqueous solution generates radicals such

254

as ·OH, Cl· and Cl2·-.7, 19 BA was selected as a radical probe compound since it reacts with ·OH,

255

Cl· and Cl2·- (rate constants given in Table S2) but does not react with free chlorine.55

256

BA degradation in a 30 mM Na2SO4 electrolyte solution at 25 mA/cm2 was observed only

257

with SbSn/CoTi/Ir* anodes (Figure 3a). When current density was increased to 50 mA/cm2, BA

258

degradation was observed with SbSn/CoTi/Ir anodes (Figure 3b), but was not observed with

259

CoTi/Ir and Ti/Ir anodes without added Sb-SnO2 islands (data not shown). BA could be degraded

260

via direct oxidation on BDD electrodes at high oxidation potentials (i.e., 2.4 VNHE).56 However,

261

this pathway was excluded in the current study as the same current responses in linear sweeping

262

voltammetry were observed in Na2SO4 in the absence or presence of 1 mM BA (data not shown).

263

The contribution of sulfate radical is excluded as the same BA decay kinetic was observed in 30

264

mM NaNO3 electrolyte (Figure S5). Thus, degradation of BA is attributed to reaction with ·OH.

265

Assuming that the generation of ·OH is a zero-order reaction, the generation rate (rHO·; M/s)

266

could be estimated by fitting BA degradation data with a kinetic model (Table 1). SbSn/CoTi/Ir*

267

was found to be more efficient than SbSn/CoTi/Ir in terms of ·OH generation (Figure 3a vs. b).

268

The steady state ·OH radical concentrations were calculated to be 2.6 × 10-15 and 1.4 × 10-15

269

mol/L for SbSn/CoTi/Ir* and SbSn/CoTi/Ir anodes, respectively.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 16 of 31

270 271

Figure 3. BA degradation with (a) SbSn/CoTi/Ir* and (b) SbSn/CoTi/Ir anodes under

272

variable current densities (L: 25 mA/cm2, H: 50 mA/cm2) and initial Cl- concentration (30

273

and 60 mM). Error bars represent the standard deviation.

274

generated by (c) SbSn/CoTi/Ir* and (d) SbSn/CoTi/Ir to BA degradation. Schematic

275

illustrations of (e) reactive species generation mechanism and (f) active site distribution.

Contributions of radical

276

BA degradation was accelerated in the presence of 30 mM NaCl (Figure 3a and b). This

277

implies that more radicals were generated in the presence of Cl-. It is well known that Cl· reacts

278

at similar rates as ·OH with organic molecules. However, Cl·, which is involved in the V-H step

279

(green line in Figure 3e), is assumed to be surface-bound and to rapidly combine with local

280

Cl· or Cl-, and hence is unlikely to contribute to BA degradation. This assertion was supported

281

by the lack of BA degradation that was observed with CoTi/Ir and Ti/Ir anodes, despite FC

282

production (data not shown). The ·OH radicals can be quenched by Cl- to generate less reactive

283

chlorine radicals (black lines in Figure 3e). A kinetic model in which ·OH is the only radical

284

species failed to simulate the observed enhanced BA degradation rate (Figure S5). Therefore,

ACS Paragon Plus Environment

Page 17 of 31

Environmental Science & Technology

285

there must be additional reactive radical inputs. One possibility involves the contribution by Sb-

286

SnO2 promoting the one-electron oxidation of Cl- to produce free Cl· (Figure 3e and f): Sb−SnO

287

2 Cl-  → Cl ⋅ + e-

288

Generation of free Cl· on SnO2 during electrolysis has been reported previously,57 although

289

without substantiating experimental evidence. The kinetic model result, however, provides a self-

290

consistent kinetic argument for the contributions of the various chlorine radical species to the

291

overall rates. The kinetic model was also used to simulate degradation of BA in both Na2SO4 and

292

NaCl electrolytes provided that eqn (12) was included.

(12)

293

The first-order rate constant for Cl· formation (kCl·; s-1) obtained by model fitting was found

294

to be more than two orders of magnitude higher than that for ·OH evolution (Table 1). The

295

higher kCl· of SbSn/CoTi/Ir compared with that of SbSn/CoTi/Ir* may be explained by the Sb-

296

SnO2 islands accepting more Cl· from the Co-TiO2 sites (Figure 3f). A similar argument was

297

used to explain the activity of RuO2-coated BDD electrodes. In this example, Cl· generated on

298

active RuO2 sites was proposed to spill over to non-active BDD sites..58

299

Even though electron transfer reactions leading to the generation of Cl· and ·OH were the

300

initial radical formation steps on anode, additional modeling of the entire set of free radical

301

concentrations (·OH, O·-, Cl·, Cl2·- and ClOH·-) showed that the dominant radical species by

302

concentration was Cl2·- (Figure S6). Model simulation further indicates that the combination of

303

Cl· and Cl- is the main pathway for Cl2·- formation. An increase in Cl- concentrations from 30 to

304

60 mM resulted in a lowering of the observed BA degradation rate (Figure 3a and b). This

305

apparent inhibition could be attributed to the role played by FC scavenging ·OH and Cl·.29, 33 The

306

decreased BA degradation rate and reduced radical concentrations observed at 60 mM NaCl

307

were successfully predicted by the kinetic modeling results (dotted line in Figure 3a and b).

ACS Paragon Plus Environment

Environmental Science & Technology

Page 18 of 31

308

The relative contributions of various free radicals to the BA degradation rate were estimated

309

by simulating the oxidation of BA by a specific target radical while excluding other radical

310

reactions in the model (Figure 3c and d). It appears that Cl· was the major contributor to BA

311

oxidation, followed by Cl2·- and ·OH, even though Cl2·- has the highest concentration. This was

312

due to the much higher reaction rate of BA with Cl· as compared to Cl2·- (1.9 × 1010 vs. 2.0 × 106

313

M-1s-1).

314

Based on the calibrated rate constants, the kinetic model is able to predict reactive species

315

formation and steady-state concentrations. In general, the FC concentrations were found to be ten

316

orders of magnitude higher than radical concentrations.

317

Wastewater electrolysis

318 319

The active SbSn/CoTi/Ir anode and non-active SbSn/CoTi/Ir* anode were tested in terms of their potential for domestic (i.e., exclusively human waste) wastewater treatment.

320 321 322 323

ACS Paragon Plus Environment

Environmental Science & Technology

SbSn/CoTi/Ir* L30

(c)

L30 50

L60 H30

8 6 L30

4

L60

2

0

1

2

Time (h)

3

4

40

20

H30

0

0

60

0

0

1

2

3

4

Time (h)

Sb H 30 Sn /C L3 o 0 Ti/I r*

100

TOC removal (%)

SbSn/CoTi/Ir* L30

150

NH4+ (mM)

10

200

COD (mg/L)

Commercial L30

(b) 12

Commercial L30

L6 0

(a) 250

L3 0

Page 19 of 31

324 325

Figure 4. Concentration vs. time profiles of (a) COD and (b) NH4+ in human wastewater

326

electrolyzed by SbSn/CoTi/Ir and SbSn/CoTi/Ir* anodes under various current densities (L:

327

25, H: 50 mA/cm2) and initial Cl- concentration (30, 60 mM). (c) Removal of TOC after 4 h

328

electrolysis. Error bars represent standard deviation. All data are collected from

329

SbSn/CoTi/Ir anode except the one labeled with SbSn/CoTi/Ir*.

330

In terms of COD removal, the SbSn/CoTi/Ir anode outperforms the commercial anode but

331

was less efficient than the SbSn/CoTi/Ir* anode at 25 mA/cm2 in 30 mM Cl- (Figure 4a).

332

Assuming that direct oxidation of COD is insignificant, then the COD removal obtained with the

333

SbSn/CoTi/Ir anode at 25 mA/cm2 should take place exclusively via FC mediated oxidation.

334

Conversely, our calculations showed that the radical-mediated oxidation pathways contributed

335

up to 80% of COD removal on SbSn/CoTi/Ir* anode (Figure S7 and Text S1). This result

336

suggests that the radicals produced by the ‘non-active’ SbSn/CoTi/Ir* anodes were more

337

efficient for COD removal than FC alone. However, TOC analysis showed that complete

338

mineralization of the organic carbon in human wastewater was not observed with the

339

SbSn/CoTi/Ir* anode (Figure 4c), in spite of complete COD removal. This may be due, in part,

340

to the contribution of Cl2·- as the dominated radical species. Cl2·- reacts with organics via

ACS Paragon Plus Environment

Environmental Science & Technology

341

hydrogen abstraction, electrophilic addition and electron transfer.19 The residual TOC may be

342

due to the formation of chlorinated byproducts or to an accumulation of formate and oxalate.

343

For example, our recent study using commercially available Ir-based anodes found that 4 h

344

human wastewater treatment via chlorine-mediated EO will form trihalomethanes and haloacetic

345

acids.59 Considering that the concentrations are generally within the range of those reported for

346

secondary effluent after disinfection process and swimming pool waters, the treated water should

347

be safe for non-potable reuse.

Page 20 of 31

348

Increasing the Cl- concentration to 60 mM significantly enhanced organic matter removal for

349

the SbSn/CoTi/Ir anode system. This is likely due to enhanced FC evolution, which effectively

350

compensated for the anode’s inability to generate radicals at 25 mA/cm2. At 50 mA/cm2,

351

SbSn/CoTi/Ir anodes produced more FC accompanied with sufficient radicals to achieve

352

complete COD removal and greater than 50% TOC removal. In this case, the contributions of

353

chlorine and radical mediated oxidation to COD removal were calculated to be 94% and 6%

354

(Figure S7 and Text S1).

355

As expected, SbSn/CoTi/Ir anodes outperformed SbSn/CoTi/Ir* and commercial anodes for

356

NH4+ removal (Figure 4b), since NH4+ removal during electrochemical treatment is achieved via

357

breakpoint chlorination,6, 7 which, in turn, is an indirect measure of the CER activity. Most NH4+

358

was converted into N2 with a smaller fraction oxidized to NO3- (Figure S8). The SbSn/CoTi/Ir

359

anode that was operated at L60 and H30 was capable of removing about 74% of total nitrogen

360

after 4 h electrolysis (Figure S8).

361

TC and FC concentrations were low (< 2 mM) during electrochemical wastewater treatment

362

(30 mM Cl-, 25 mA/cm2) with SbSn/CoTi/Ir* and SbSn/CoTi/Ir anodes. Chorine was consumed

363

during breakpoint chlorination or by wastewater organic matter degradation within the electrical

ACS Paragon Plus Environment

Page 21 of 31

Environmental Science & Technology

364

double layer, and thus was unable to diffuse into the bulk solution phase. Generation of both TC

365

and FC was only observed with SbSn/CoTi/Ir anodes after complete NH4+ and COD removal

366

(Figure S9). Up to 5 mM ClO3- was produced after 4 h electrolysis (30 mM Cl-; 25 mA/cm2)

367

with both types of anodes (Figure S9). Increasing either current or the Cl- concentration

368

increased ClO3- production (i.e., 10-11 mM maximum). These trends were in agreement with

369

experiments in NaCl solutions, although concentrations were about 50% lower.

370

Significant concentrations of ClO4- (6 mM) were produced by SbSn/CoTi/Ir* anode after 4 h

371

electrolysis of wastewater under low Cl- and low current conditions (i.e., 30 mM Cl-; 25 mA/cm2;

372

Figure S9). However, SbSn/CoTi/Ir anodes only formed relatively low concentrations of ClO4-

373

(0.85 mM) under high current conditions (i.e., 30 mM Cl-; 50 mA/cm2).

374

The SbSn/CoTi/Ir anode, which was operated at 60 mM Cl- and 25 mA/cm2, consumed less

375

energy than the SbSn/CoTi/Ir* anode for COD and NH4+ removal (370 kWh/kg COD; 383

376

kWh/kg NH4+). These values are still higher than those reported in the EO of leachate and

377

reverse osmosis concentrate,11, 60 probably due to the lower conductivity (3.2 mS/cm) of human

378

wastewater. Reducing the electrode spacing or increasing the wastewater conductivity may be

379

able to further lower the energy consumption.

380

In conclusion, the SbSn/CoTi/Ir anode was observed to be the most effective anode in terms

381

of durability, reactive species generation, pollutant removal, by-product formation, and energy

382

consumption. More efficient wastewater treatment provided by an active anode (SbSn/CoTi/Ir)

383

as compared to a non-active anode (SbSn/CoTi/Ir*) highlights the limitation of non-active anodes

384

for wastewater treatment due to ·OH quenching by Cl- and FC. In addition, non-active anodes

385

produce a significant amount of ClO4-. Under appropriate conditions, wastewater electrolysis

ACS Paragon Plus Environment

Environmental Science & Technology

Page 22 of 31

386

mediated by electrochemically produced FC may be able to outperform radical-assisted

387

electrolysis.

388

From an engineering point of view, wastewater treated in an appropriately designed reactor

389

equipped with SbSn/CoTi/Ir anodes should be suitable for non-potable water reuse (e.g., as

390

recycled toilet flushing water based on color (Figure S10) and COD removal), as well as for

391

disinfection, which is provided by the residual FC. The semiconductor electrolytic reactors can

392

be easily automated (e.g., reaching the breakpoint for NH4+ chlorination can be used as a signal

393

to end batch treatment). They should be an excellent fit for use in decentralized wastewater

394

treatment.

395

ASSOCIATED CONTENT

396

Figures provided in supporting information include SEM, LSV, and ECSA measurement of

397

anodes, model simulation results, time profiles of ions, the calculation of FC contribution to

398

pollutant removal and energy consumption in wastewater electrolysis. Tables include human

399

wastewater composition and details of the kinetic model.

400

ACKNOWLEDGEMENT

401 402

The authors gratefully acknowledge the financial support of Bill and Melinda Gates Foundation (BMGF-RTTC Grant, OPP1111246).

ACS Paragon Plus Environment

Page 23 of 31

Environmental Science & Technology

References 1. Levine, A. D.; Asano, T. Peer Reviewed: Recovering Sustainable Water from Wastewater. Environ. Sci. Technology 2004, 38 (11), 201A-208A. 2. Martínez-Huitle, C. A.; Rodrigo, M. A.; Sirés, I.; Scialdone, O. Single and Coupled Electrochemical Processes and Reactors for the Abatement of Organic Water Pollutants: A critical review. Chem. Rev. 2015, 115 (24), 13362-13407. 3. Canizares, P.; Paz, R.; Sáez, C.; Rodrigo, M. A. Costs of the Electrochemical Oxidation of Wastewaters: A Comparison with Ozonation and Fenton Oxidation Processes. J. Environ. Manage. 2009, 90 (1), 410-420. 4. Radjenovic, J.; Sedlak, D. L. Challenges and Opportunities for Electrochemical Processes as Next-Generation Technologies for the Treatment of Contaminated Water. Environ. Sci. Technol. 2015, 49 (19), 11292-11302. 5. Barazesh, J. M.; Hennebel, T.; Jasper, J. T.; Sedlak, D. L. Modular Advanced Oxidation Process Enabled by Cathodic Hydrogen Peroxide Production. Environ. Sci. Technol. 2015, 49 (12), 7391-7399. 6. Cho, K.; Kwon, D.; Hoffmann, M. R. Electrochemical Treatment of Human Waste Coupled with Molecular Hydrogen Production. RSC Adv. 2014, 4 (9), 4596-4608. 7. Cho, K.; Qu, Y.; Kwon, D.; Zhang, H.; Cid, C. m. A.; Aryanfar, A.; Hoffmann, M. R. Effects of Anodic Potential and Chloride Ion on Overall Reactivity in Electrochemical Reactors Designed for Solar-Powered Wastewater Treatment. Environ. Sci. Technol. 2014, 48 (4), 2377-2384.

ACS Paragon Plus Environment

Environmental Science & Technology

8. Bagastyo, A. Y.; Batstone, D. J.; Rabaey, K.; Radjenovic, J. Electrochemical Oxidation of Electrodialysed Reverse Osmosis Concentrate on Ti/Pt-IrO2, Ti/SnO2-Sb and Borondoped Diamond Electrodes. Wat. Res. 2013, 47 (1), 242-250. 9. Lin, H.; Niu, J.; Xu, J.; Huang, H.; Li, D.; Yue, Z.; Feng, C. Highly Efficient and Mild Electrochemical Mineralization of Long-Chain Perfluorocarboxylic Acids (C9-C10) by Ti/SnO2-Sb-Ce, Ti/SnO2-Sb/Ce-PbO2, and Ti/BDD Electrodes. Environ. Sci. Technol. 2013, 47 (22), 13039-13046. 10. Yang, S. Y.; Kim, D.; Park, H. Shift of the Reactive Species in the Sb-SnO2Electrocatalyzed Inactivation of E. coli and Degradation of Phenol: Effects of Nickel Doping and Electrolytes. Environ. Sci. Technol. 2014, 48 (5), 2877-2884. 11. Bagastyo, A. Y.; Batstone, D. J.; Kristiana, I.; Gernjak, W.; Joll, C.; Radjenovic, J. Electrochemical Oxidation of Reverse Osmosis Concentrate on Boron-Doped Diamond Anodes at Circumneutral and Acidic pH. Wat. Res. 2012, 46 (18), 6104-6112. 12. Dominguez-Ramos, A.; Irabien, A. Analysis and Modeling of the Continuous Electrooxidation Process for Organic Matter Removal in Urban Wastewater Treatment. Ind. Eng. Chem. Res. 2013, 52 (22), 7534-7540. 13. Zöllig, H.; Remmele, A.; Fritzsche, C.; Morgenroth, E.; Udert, K. Formation of Chlorination Byproducts and Their Emission Pathways in Chlorine Mediated Electrooxidation of Urine on Active and Nonactive Type Anodes. Environ. Sci. Technol. 2015, 49 (18), 1062-11069. 14. Trasatti, S. Electrocatalysis in the Anodic Evolution of Oxygen and Chlorine. Electrochim. Acta 1984, 29 (11), 1503-1512.

ACS Paragon Plus Environment

Page 24 of 31

Page 25 of 31

Environmental Science & Technology

15. Zhu, X. P.; Tong, M. P.; Shi, S. Y.; Zhao, H. Z.; Ni, J. R. Essential Explanation of the Strong Mineralization Performance of Boron-Doped Diamond Electrodes. Environ. Sci. Technol. 2008, 42 (13), 4914-4920. 16. Farhat, A.; Keller, J.; Tait, S.; Radjenovic, J. Removal of Persistent Organic Contaminants by Electrochemically Activated Sulfate. Environ. Sci. Technol. 2015, 49 (24), 14326-14333. 17. Malpass, G.; Miwa, D.; Mortari, D.; Machado, S.; Motheo, A. Decolorisation of Real Textile Waste Using Electrochemical Techniques: Effect of the Chloride Concentration. Wat. Res. 2007, 41 (13), 2969-2977. 18. Bonfatti, F.; Ferro, S.; Lavezzo, F.; Malacarne, M.; Lodi, G.; De Battisti, A. Electrochemical Incineration of Glucose as a Model Organic Substrate. II. Role of Active Chlorine Mediation. J. Electrochem. Soc. 2000, 147 (2), 592-596. 19. Park, H.; Vecitis, C. D.; Hoffmann, M. R. Electrochemical Water Splitting Coupled with Organic Compound Oxidation: The role of Active Chlorine Species. J. Phys. Chem. B 2009, 113 (18), 7935-7945. 20. Cho, K.; Hoffmann, M. R. Urea Degradation by Electrochemically Generated Reactive Chlorine Species: Products and Reaction Pathways. Environ. Sci. Technol. 2014, 48 (19), 11504-11511. 21. Cho, K.; Hoffmann, M. R. BixTi1-xOz Functionalized Heterojunction Anode with an Enhanced Reactive Chlorine Generation Efficiency in Dilute Aqueous Solutions. Chem. Mater. 2015, 27 (6), 2224-2233.

ACS Paragon Plus Environment

Environmental Science & Technology

22. McCrory, C. C.; Jung, S.; Peters, J. C.; Jaramillo, T. F. Benchmarking Heterogeneous Electrocatalysts for the Oxygen Evolution Reaction. Journal of the American Chemical Society 2013, 135 (45), 16977-16987. 23. Hoffmann, M. R.; Cho, K.; Cid, C.; Yan, Q. Development of a Self-Contained, PVPowered Domestic Toilet and Electrochemical Wastewater Treatment System Suitable for the Developing World. International Conference on Sustainable Energy & Environmental Sciences (SEES). Proceedings, 2014; Global Science and Technology Forum: 2014; p 34. 24. Ianni, J. C. Kintecus, Windows Version 5.75. http://www.kintecus.com (accessed December 1, 2015) 25. Matthew, B.; Anastasio, C. A Chemical Probe Technique for the Determination of Reactive Halogen Species in Aqueous Solution: Part 1-Bromide Solutions. Atmos. Chem. Phys. 2006, 6 (9), 2423-2437. 26. Wang, T. X.; Margerum, D. W. Kinetics of Reversible Chlorine Hydrolysis: Temperature Dependence and General-Acid/Base-Assisted Mechanisms. Inorg. Chem. 1994, 33 (6), 1050-1055. 27. Tao, L.; Han, J.; Tao, F.-M. Correlations and Predictions of Carboxylic Acid pKa Values Using Intermolecular Structure and Properties of Hydrogen-Bonded Complexes. J. Phys. Chem. A 2008, 112 (4), 775-782. 28. Buxton, G. V.; Greenstock, C. L.; Helman, W. P.; Ross, A. B. Critical Review of Rate Constants for Reactions of Hydrated Electrons, Hydrogen Atoms and Hydroxyl Radicals ( OH/ O-) in Aqueous Solution. J. Phys. Chem. Ref. Data 1988, 17 (2), 513-886.

ACS Paragon Plus Environment

Page 26 of 31

Page 27 of 31

Environmental Science & Technology

29. Jayson, G.; Parsons, B.; Swallow, A. J. Some Simple, Highly Reactive, Inorganic Chlorine Derivatives in Aqueous Solution. Their Formation Using Pulses of Radiation and Their Role in the Mechanism of the Fricke Dosimeter. J. Chem. Soc., Faraday Trans. 1 1973, 69, 1597-1607. 30. Kläning, U. K.; Wolff, T. Laser Flash Photolysis of HCIO, CIO-, HBrO, and BrO- in Aqueous Solution. Reactions of Cl

and Br

Atoms. Beri. Bunsenges. Phys. Chem.

1985, 89 (3), 243-245. 31. Grebel, J. E.; Pignatello, J. J.; Mitch, W. A. Effect of Halide Ions and Carbonates on Organic Contaminant Degradation by Hydroxyl Radical-Based Advanced Oxidation Processes in Saline Waters. Environ. Sci. Technol 2010, 44 (17), 6822-6828. 32. Wu, D.; Wong, D.; Di Bartolo, B. Evolution of Cl2- in Aqueous NaCl Solutions. J. Photochem. 1980, 14 (4), 303-310. 33. Connick, R. E. The Interaction of Hydrogen Peroxide and Hypochlorous Acid in Acidic Solutions Containing Chloride Ion. J. Am. Chem. Soc. 1947, 69 (6), 1509-1514. 34. Zehavi, D.; Rabani, J. Oxidation of Aqueous Bromide Ions by Hydroxyl Radicals. Pulse Radiolytic Investigation. J. Phys. Chem. A 1972, 76 (3), 312-319. 35. Mártire, D. O.; Rosso, J. A.; Bertolotti, S.; Le Roux, G. C.; Braun, A. M.; Gonzalez, M. C. Kinetic Study of the Reactions of Chlorine Atoms and Cl2-Radical Anions in Aqueous Solutions. II. Toluene, Benzoic Acid, and Chlorobenzene. J. Phys. Chem. A 2001, 105 (22), 5385-5392. 36. Hasegawa, K.; Neta, P. Rate Constants and Mechanisms of Reaction of Chloride (Cl2-) Radicals. J. Phys. Chem. 1978, 82 (8), 854-857.

ACS Paragon Plus Environment

Environmental Science & Technology

37. Correa-Lozano, B.; Comninellis, C.; De Battisti, A. Service life of Ti/SnO2–Sb2O5 anodes. J. Appl. Electrochem. 1997, 27 (8), 970-974. 38. Lettenmeier, P.; Wang, L.; Golla Schindler, U.; Gazdzicki, P.; Cañas, N. A.; Handl, M.; Hiesgen, R.; Hosseiny, S. S.; Gago, A. S.; Friedrich, K. A. Nanosized IrOx-Ir Catalyst with Relevant Activity for Anodes of Proton Exchange Membrane Electrolysis Produced by a Cost

Effective Procedure. Angew. Chem. 2015, 128 (2), 752-756.

39. Minguzzi, A.; Locatelli, C.; Lugaresi, O.; Achilli, E.; Cappelletti, G.; Scavini, M.; Coduri, M.; Masala, P.; Sacchi, B.; Vertova, A. Easy Accommodation of Different Oxidation States in Iridium Oxide Nanoparticles With Different Hydration Degree as Water Oxidation Electrocatalysts. ACS Catal. 2015. 40. Imanishi, A.; Tsuji, E.; Nakato, Y. Dependence of the Work Function of TiO2 (rutile) on Crystal Faces, Studied by a Scanning Auger Microprobe. J. Phys. Chem. C 2007, 111 (5), 2128-2132. 41. Riga, J.; Tenret-Noel, C.; Pireaux, J.; Caudano, R.; Verbist, J.; Gobillon, Y. Electronic Structure of Rutile Oxides TiO2, RuO2 and IrO2 Studied by X-ray Photoelectron Spectroscopy. Phys. Scrip. 1977, 16 (5-6), 351. 42. Yang, Y.; Zhang, S.; Wang, S.; Zhang, K.; Wang, H.; Huang, J.; Deng, S.; Wang, B.; Wang, Y.; Yu, G. Ball Milling Synthesized MnO x as Highly Active Catalyst for Gaseous POPs Removal: Significance of Mechanochemically Induced Oxygen Vacancies. Environ. Sci. Technol. 2015, 49 (7), 4473-4480. 43. Reier, T.; Oezaslan, M.; Strasser, P. Electrocatalytic Oxygen Evolution Reaction (OER) on Ru, Ir, and Pt catalysts: A Comparative Study of Nanoparticles and Bulk Materials. ACS Catal. 2012, 2 (8), 1765-1772.

ACS Paragon Plus Environment

Page 28 of 31

Page 29 of 31

Environmental Science & Technology

44. Smith, R. D.; Sporinova, B.; Fagan, R. D.; Trudel, S.; Berlinguette, C. P. Facile Photochemical Preparation of Amorphous Iridium Oxide Films for Water Oxidation Catalysis. Chem. Mater. 2014. 45. Consonni, V.; Trasatti, S.; Pollak, F.; O'Grady, W. Mechanism of Chlorine Evolution on Oxide Anodes Study of pH Effects. J. Electroanal. Chem. Interfacial Electrochem. 1987, 228 (1), 393-406. 46. Trasatti, S. Progress in the Understanding of the Mechanism of Chlorine Evolution at Oxide Electrodes. Electrochim. Acta 1987, 32 (3), 369-382. 47. Hansen, H. A.; Man, I. C.; Studt, F.; Abild-Pedersen, F.; Bligaard, T.; Rossmeisl, J. Electrochemical Chlorine Evolution at Rutile Oxide (110) Surfaces. Phys. Chem. Chem. Phys. 2010, 12 (1), 283-290. 48. Exner, K. S.; Anton, J.; Jacob, T.; Over, H. Controlling Selectivity in the Chlorine Evolution Reaction over RuO2 Based Catalysts. Angew. Chem. Inter. Ed. 2014, 53 (41), 11032-11035. 49. Tilak, B. Kinetics of Chlorine Evolution-A Comparative Study. J. Electrochem. Soc. 1979, 126 (8), 1343-1348. 50. Azizi, O.; Hubler, D.; Schrader, G.; Farrell, J.; Chaplin, B. P. Mechanism of perchlorate formation on boron-doped diamond film anodes. Environ. Sci. Technol. 2011, 45 (24), 10582-10590. 51. Bergmann, M. H.; Rollin, J.; Iourtchouk, T. The Occurrence of Perchlorate During Drinking Water Electrolysis using BDD Anodes. Electrochim. Acta 2009, 54 (7), 21022107.

ACS Paragon Plus Environment

Environmental Science & Technology

Page 30 of 31

52. Tasaka, A.; Tojo, T. Anodic Oxidation Mechanism of Hypochlorite Ion on Platinum Electrode in Alkaline Solution. J. Electrochem. Soc. 1985, 132 (8), 1855-1859. 53. Jung, Y. J.; Baek, K. W.; Oh, B. S.; Kang, J.-W. An Investigation of the Formation of Chlorate and Perchlorate During Electrolysis Using Pt/Ti Electrodes: The Effects of pH and Reactive Oxygen Species and the Results of Kinetic Studies. Wat. Res. 2010, 44 (18), 5345-5355. 54. Bard, A. J.; Faulkner, L. R. Electrochemical Methods: Fundamentals and Applications. 2 ed.; Wiley New York: 2001. 55. Fang, J.; Fu, Y.; Shang, C. The Roles of Reactive Species in Micropollutant Degradation in the UV/free Chlorine System. Environ. Sci. Technol 2014, 48 (3), 1859-1868. 56. Montilla, F.; Michaud, P.; Morallon, E.; Vazquez, J.; Comninellis, C. Electrochemical Oxidation of Benzoic Acid at Boron-Doped Diamond Electrodes. Electrochim. Acta 2002, 47 (21), 3509-3513. 57. Wu, W.; Huang, Z.-H.; Lim, T.-T. Enhanced Electrochemical Oxidation of Phenol Using a

Hydrophobic

TiO2-NTs/SnO2-Sb-PTFE

Electrode

Prepared

by

Pulse

Electrodeposition. RSC Adv. 2015, 5 (41), 32245-32255. 58. Ferro, S.; De Battisti, A. Electrocatalysis and Chlorine Evolution Reaction at Ruthenium Dioxide Deposited on Conductive Diamond. J. Phys. Chem. B 2002, 106 (9), 2249-2254. 59. Huang, X.; Qu, Y.; Cid, C. A.; Finke, C.; Hoffmann, M. R.; Lim, K.; Jiang, S. C. Electrochemical Disinfection of Toilet Wastewater Using Wastewater Electrolysis Cell. Wat. Res. 2016, 92, 164-172.

ACS Paragon Plus Environment

Page 31 of 31

Environmental Science & Technology

60. Anglada, A.; Ortiz, D.; Urtiaga, A.; Ortiz, I. Electrochemical Oxidation of Landfill Leachates at Pilot Scale: Evaluation of Energy Needs. Wat. Sci. Technol. 2010, 61 (9), 2211-2217.

ACS Paragon Plus Environment